vivianite as an important iron phosphate precipitate in sewage … · accepted manuscript vivianite...

58
Delft University of Technology Vivianite as an important iron phosphate precipitate in sewage treatment plants Wilfert, P.; Mandalidis, A.; Dugulan, A. I.; Goubitz, K.; Korving, L.; Temmink, H.; Witkamp, G. J.; Van Loosdrecht, M. C M DOI 10.1016/j.watres.2016.08.032 Publication date 2016 Document Version Accepted author manuscript Published in Water Research Citation (APA) Wilfert, P., Mandalidis, A., Dugulan, A. I., Goubitz, K., Korving, L., Temmink, H., Witkamp, G. J., & Van Loosdrecht, M. C. M. (2016). Vivianite as an important iron phosphate precipitate in sewage treatment plants. Water Research, 104, 449-460. https://doi.org/10.1016/j.watres.2016.08.032 Important note To cite this publication, please use the final published version (if applicable). Please check the document version above. Copyright Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim. This work is downloaded from Delft University of Technology. For technical reasons the number of authors shown on this cover page is limited to a maximum of 10.

Upload: others

Post on 13-Feb-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • Delft University of Technology

    Vivianite as an important iron phosphate precipitate in sewage treatment plants

    Wilfert, P.; Mandalidis, A.; Dugulan, A. I.; Goubitz, K.; Korving, L.; Temmink, H.; Witkamp, G. J.; VanLoosdrecht, M. C MDOI10.1016/j.watres.2016.08.032Publication date2016Document VersionAccepted author manuscriptPublished inWater Research

    Citation (APA)Wilfert, P., Mandalidis, A., Dugulan, A. I., Goubitz, K., Korving, L., Temmink, H., Witkamp, G. J., & VanLoosdrecht, M. C. M. (2016). Vivianite as an important iron phosphate precipitate in sewage treatmentplants. Water Research, 104, 449-460. https://doi.org/10.1016/j.watres.2016.08.032

    Important noteTo cite this publication, please use the final published version (if applicable).Please check the document version above.

    CopyrightOther than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consentof the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons.

    Takedown policyPlease contact us and provide details if you believe this document breaches copyrights.We will remove access to the work immediately and investigate your claim.

    This work is downloaded from Delft University of Technology.For technical reasons the number of authors shown on this cover page is limited to a maximum of 10.

    https://doi.org/10.1016/j.watres.2016.08.032https://doi.org/10.1016/j.watres.2016.08.032

  • Accepted Manuscript

    Vivianite as an important iron phosphate precipitate in sewage treatment plants

    P. Wilfert, A. Mandalidis, A.I. Dugulan, K. Goubitz, L. Korving, H. Temmink, G.J.

    Witkamp, M.C.M. Van Loosdrecht

    PII: S0043-1354(16)30632-7

    DOI: 10.1016/j.watres.2016.08.032

    Reference: WR 12303

    To appear in: Water Research

    Received Date: 23 May 2016

    Revised Date: 8 August 2016

    Accepted Date: 18 August 2016

    Please cite this article as: Wilfert, P., Mandalidis, A., Dugulan, A.I., Goubitz, K., Korving, L., Temmink,

    H., Witkamp, G.J., Van Loosdrecht, M.C.M., Vivianite as an important iron phosphate precipitate in

    sewage treatment plants, Water Research (2016), doi: 10.1016/j.watres.2016.08.032.

    This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to

    our customers we are providing this early version of the manuscript. The manuscript will undergo

    copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please

    note that during the production process errors may be discovered which could affect the content, and all

    legal disclaimers that apply to the journal pertain.

    http://dx.doi.org/10.1016/j.watres.2016.08.032

  • ACCEPTED MANUSCRIPT

    1

    1 Vivianite as an important iron phosphate precipitate in sewage

    2 treatment plants

    3 P. Wilferta,b

    , A. Mandalidisa, A.I. Dugulan

    c, K. Goubitz

    c, L. Korving

    a,*, H. Temmink

    a,d, G.J.

    4 Witkampb and M.C.M Van Loosdrecht

    b

    5 aWetsus, European Centre Of Excellence for Sustainable Water Technology, Oostergoweg 7,

    6 8911 MA, Leeuwarden, The Netherlands

    7 bDept. Biotechnology, Delft Univ Technol, Van der Maasweg 9, 2629 HZ Delft, The Netherlands

    8 cFundamental Aspects Mat & Energy Grp, Delft Univ Technol, Mekelweg 15, 2629 JB Delft, The

    9 Netherlands

    10 dSub-department of Environmental Technology, Wageningen University, P.O. Box 8129, 6700

    11 EV Wageningen, The Netherlands

    12 *Corresponding author: Phone: +31-58-2843160; e-mail: [email protected]

    13 Abbreviations

    14 CPR – Chemical Phosphorus Removal

    15 COD – Chemical Oxygen Demand

    16 DO - Dissolved Oxygen

    17 DOC - Dissolved Organic Carbon

    18 EBPR – Enhanced Biological Phosphorus Removal

    19 Fe2O3 – Hematite

    20 FeP - Iron Phosphorus Compounds

    mailto:[email protected]

  • ACCEPTED MANUSCRIPT

    2

    21 FeSx - Iron Sulphide Compounds

    22 IRB - Iron Reducing Bacteria

    23 o-P - Orthophosphate

    24 SRT – Solid Retention Time

    25 STP – Sewage Treatment Plant

    26 TS - Total Solids

    27 Abstract

    28 Iron is an important element for modern sewage treatment, inter alia to remove phosphorus from

    29 sewage. However, phosphorus recovery from iron phosphorus containing sewage sludge, without

    30 incineration, is not yet economical. We believe, increasing the knowledge about iron-phosphorus

    31 speciation in sewage sludge can help to identify new routes for phosphorus recovery. Surplus and

    32 digested sludge of two sewage treatment plants was investigated. The plants relied either solely on

    33 iron based phosphorus removal or on biological phosphorus removal supported by iron dosing.

    34 Mössbauer spectroscopy showed that vivianite and pyrite were the dominating iron compounds in

    35 the surplus and anaerobically digested sludge solids in both plants. Mössbauer spectroscopy and

    36 XRD suggested that vivianite bound phosphorus made up between 10 and 30 % (in the plant

    37 relying mainly on biological removal) and between 40 and 50 % of total phosphorus (in the plant

    38 that relies on iron based phosphorus removal). Furthermore, Mössbauer spectroscopy indicated

    39 that none of the samples contained a significant amount of Fe(III), even though aerated treatment

    40 stages existed and although besides Fe(II) also Fe(III) was dosed. We hypothesize that

    41 chemical/microbial Fe(III) reduction in the treatment lines is relatively quick and triggers

    42 vivianite formation. Once formed, vivianite may endure oxygenated treatment zones due to slow

    43 oxidation kinetics and due to oxygen diffusion limitations into sludge flocs. These results indicate

  • ACCEPTED MANUSCRIPT

    3

    44 that vivianite is the major iron phosphorus compound in sewage treatment plants with moderate

    45 iron dosing. We hypothesize that vivianite is dominating in most plants where iron is dosed for

    46 phosphorus removal which could offer new routes for phosphorus recovery.

    47 Keywords:

    48 Iron, Phosphorus, Sewage, Sewage sludge, Mössbauer spectroscopy, Vivianite

    49 1 Introduction

    50 Phosphorus (P) is an essential element for all life. It is often a limiting nutrient for crops and thus

    51 a crucial part of fertilizers. Currently, the use of P is not sustainable and its supply is not

    52 guaranteed in the future: (I) Phosphate rock reservoirs, the main source of P for fertilizers, are

    53 depleting (Scholz and Wellmer, 2016; Walan et al., 2014), (II) these reservoirs are located in a

    54 few countries (De Ridder et al., 2012), (III) for current P applications and depletions regional

    55 imbalances exist (Macdonald et al., 2011; van Dijk et al., 2016) and (IV) P surpluses cause

    56 eutrophication in surface waters (Carpenter, 2008). The recovery of P from secondary resources

    57 would help to make its use in our society circular and more sustainable (Carpenter and Bennett,

    58 2011; Childers et al., 2011).

    59 Sewage is an important secondary source for P (van Dijk et al., 2016). In sewage treatment plants

    60 (STPs), P is typically removed to diminish eutrophication in surface waters by chemical P

    61 removal (CPR) or enhanced biological P removal (EPBR). In both cases, P is concentrated in the

    62 sewage sludge. Iron (Fe) dosing for CPR is efficient, simple and cheap (Geraarts et al., 2007; Paul

    63 et al., 2001; WEF, 2011). Future energy producing STPs rely on chemical P and chemical oxygen

    64 demand (COD) removal (Böhnke, 1977; Wilfert et al., 2015). Additionally, Fe is commonly

  • ACCEPTED MANUSCRIPT

    4

    65 applied in modern sewage treatment also for other reasons than CPR. Ferric, Fe(III) and ferrous,

    66 Fe(II) iron salts are dosed as flocculants to remove COD (Li, 2005), to prevent the emission of

    67 hydrogen sulphide (H2S) in sewer systems and digesters (Hvitved-Jacobsen et al., 2013; Nielsen et

    68 al., 2005) and to improve sludge dewatering (Higgins and Murthy, 2006). Additionally, Fe may

    69 originate from groundwater intrusion into the sewer systems (Hvitved-Jacobsen et al., 2013;

    70 Kracht et al., 2007). Thus, in most STPs, part of the P will be bound to Fe. An economic feasible

    71 recovery of P from sewage sludge containing iron phosphorus compounds (FeP), without sludge

    72 incineration, is a technological challenge that remains unsolved, also due to scarce information on

    73 FeP mineralogy in STPs (Wilfert et al 2015).

    74 The initial reactions, after Fe(III) or Fe(II) addition to sewage and the subsequent removal of P are

    75 complex (El Samrani et al., 2004; Luedecke et al., 1989; Smith et al., 2008; Takács et al., 2006).

    76 These reactions are important as they drive primary P removal from sewage by bringing P from

    77 the liquid to the solid phase. In STPs, the solid retention time (SRT) can be a few hours, as in the

    78 A-stage of AB-processes (Böhnke, 1977; Böhnke et al., 1997), but it is usually on the time scale

    79 of 5-20 days when the conventional activated sludge process is applied (Tchobanoglous et al.,

    80 2013). In those processes, alternating redox conditions are applied to achieve COD and nitrogen

    81 removal. Hence, once formed the initial FeP may change due to oxidation of Fe(II) or reduction of

    82 Fe(III) respectively (Nielsen, 1996; Nielsen et al., 2005; Nielsen and Nielsen, 1998; Rasmussen

    83 and Nielsen, 1996) or due to aging effects (Recht and Ghassemi, 1970; Szabó et al., 2008). Most

    84 likely, the FeP, that end up in the surplus sludge and that determine the P removal efficiency of

    85 STPs, differ from the initial precipitates.

    86 Several researchers reported the ferrous iron phosphate mineral vivianite (Fe(II)3[PO4]2· 8H2O) in

    87 surplus sludge and anaerobically digested sludge (Frossard et al., 1997; Ghassemi and Recht,

  • ACCEPTED MANUSCRIPT

    5

    88 1971; Seitz et al., 1973; Singer, 1972). Frossard et al., 1997 were able to quantify vivianite in

    89 sewage sludge using Mössbauer spectroscopy even though the sludge samples in this study were

    90 exposed to air. This could have resulted in full/partial oxidation of Fe(II) compounds and partial

    91 transformation of vivianite to amorphous FeP (Roldan et al., 2002) or to other changes of the P

    92 fractions (Kraal et al., 2009). Additionally, all Mössbauer measurements were done at room

    93 temperature (300 K). Complex samples should ideally be measured at lower temperatures (e.g.

    94 4.2 K) as well, to reveal unambiguously the spectral contributions and magnetic properties of the

    95 Fe phases. (Murad and Cashion, 2004).

    96 We have investigated two STPs, with different treatment strategies to determine the fate of Fe and

    97 FeP during treatment. The STP Leeuwarden applies EBPR, additionally respectively Fe(II) or

    98 Fe(III) are dosed in two different treatment lines. The STP Nieuwveer uses the AB technology

    99 (Böhnke et al., 1997; De Graaff et al., 2015), here Fe(II) is dosed. AB-plants in combination with

    100 cold anammox have the potential to be energy factories (Jetten et al., 1997; Siegrist et al., 2008).

    101 The fate of Fe and FeP was evaluated by various measurements on the liquid and solid fractions of

    102 the sewage (sludge) at different locations in the treatment line. Mössbauer spectroscopy

    103 (qualitative and quantitative analyses of Fe compounds), XRD (semi-quantitative analyses of all

    104 crystalline material) and SEM-EDX (particle morphology and elemental composition) were used

    105 to characterize the solid fractions. Mass balances for P and Fe helped to identify the significance

    106 of different sources (influent, external sludge, Fe dosing) and sinks (effluent, sludge disposal) for

    107 these elements. Mössbauer spectroscopy and XRD were used to estimate P bound in vivianite and

    108 sulphide extraction was used to quantify P bound to Fe. Thereby, the P recovery potential of a

    109 technology that targets specifically on FeP was determined.

  • ACCEPTED MANUSCRIPT

    6

    110 Identifying the forms of FeP in activated sludge would help to obtain thermodynamic (e.g.

    111 equilibrium concentrations) and stoichiometric (molar Fe:P ratios) information that is necessary to

    112 develop technologies to recover P from FeP. Although, in literature, some indications for vivianite

    113 formation as major P compound during sewage treatment can be found, the role of vivianite and

    114 its importance has been neglected, the reason why this study was carried out.

    115 2 Methods & Material

    116 2.1 STPs and sampling

    117 In the AB plant Nieuwveer (influent: 75706 m3 d

    -1 in 2014), Fe(II) is added in the aerated

    118 (≈0.3 mg dissolved oxygen (DO) L-1

    ) A-stage for P and COD removal. SRTs are 15 hours in the

    119 A-stage, 16 days in the B-stage (DO in aerated sections ≈1.8 mg DO L-1

    ) and 25 days during

    120 anaerobic digestion. In the EBPR plant Leeuwarden (38,000 m3

    d-1

    in 2014), the influent is split in

    121 two treatment lines (60 % of the sewage goes to Line 1). Besides for CPR, Fe is dosed to prevent

    122 H2S emissions into the biogas during anaerobic digestion. In Line 1, Fe(III) is dosed and in Line 2

    123 Fe(II) is dosed in the nitrification zone (≈1.5 mg DO L-1

    ). SRTs before digestion are around 15

    124 days (50 % in the aerated zone) and during anaerobic digestion around 42 days. The digesters of

    125 both STPs, receive external sludge which accounts for about 30 % (Nieuwveer) and about 25 %

    126 (Leeuwarden) of the total digested sludge. At both locations, samples were taken to analyse the

    127 composition of the of the sewage (sludge). From these measurements (Table A. 1 and Table A. 2)

    128 Fe and P mass balances were calculated. To calculate Fe and P loads, average daily flow rates of

    129 the sampling days were used. Samples were taken between December 2014 and March 2015, after

    130 a period of 48 h without precipitation. The STP Leeuwarden was sampled three times. Results

  • ACCEPTED MANUSCRIPT

    7

    131 reported in Table A. 1 and for the mass balances are average values of the triplicate measurements

    132 and of the daily loads of these samplings. The STP Nieuwveer was sampled once in March 2015.

    133 Samples were stored and transported in cooling boxes on ice to reduce microbial activity. Sample

    134 processing started 1 h (Leeuwarden) and 3 h (Nieuwveer) after sampling. Sample drying started

    135 latest 8 h after sampling and was completed within 24 h. Sampling and sample processing were

    136 done under anaerobic conditions to prevent oxidation and degassing of samples. Sewage was

    137 collected; using syringes with attached tubing that were washed several times with sewage.

    138 Sewage sludge was taken from valves using a funnel with attached tubing. Samples were then

    139 filled in serum bottles. To rinse bottles, about three times their volume was flowed through the

    140 bottle by inserting the end of the tubing to the bottom of the bottle. Then the bottles were sealed

    141 with butyl rubber stoppers (referred to as anaerobic samples hereafter). With these samples the

    142 composition of the liquid phase was determined and material for solid analyses was obtained. For

    143 total solids (TS), volatile solids (VS) and to determine the total elemental composition of the

    144 sewage (sludge), separate samples, without special pre-cautions to prevent sample oxidation, were

    145 taken (hereafter, referred to as mass balance (MB) samples). Separate sampling was considered to

    146 be necessary as the TS content of the anaerobic samples could change due to rinsing of serum

    147 bottles.

    148 In Nieuwveer, the influent sample was a mixture of raw influent and recirculated effluent (40 % of

    149 the effluent is recirculated). For the mass balances, the Fe and P concentrations of the raw effluent

    150 were calculated. The external sludge sample was taken from a pre-storage tank, and contained an

    151 unknown mixture of external sludge. In Leeuwarden, P loads from external sludge were below the

    152 detection limit and had to be calculated from the difference in P loads before and after digestion.

  • ACCEPTED MANUSCRIPT

    8

    153 2.2 Analyses

    154 Oxidation-Reduction Potential (ORP), pH and conductivity were measured potentiometrically in

    155 the plants. Total elemental composition of MB samples were determined after microwave assisted

    156 acid digestion (HNO3=69%, 15 min, 180 °C) followed by ICP-OES. Total solids and VS were

    157 measured according to standard methods (APHA, AWWA, WEF, 1998). For total alkalinity

    158 measurements, 10 mL MB sample was titrated to pH=4.5 with 0.1 N HCl (APHA, AWWA, WEF,

    159 1998).

    160 The anaerobic samples were transferred into plastic centrifuge tubes inside an anaerobic glovebox

    161 (95 % N2/5 % H2, O2

  • ACCEPTED MANUSCRIPT

    9

    175 measurements, samples were filtered inside the glovebox into a zinc acetate solution (0.8 M),

    176 stored in the dark and measured after 24 h by the methylene blue method (Cline, 1969).

    177 Solid material was derived from centrifuge pellets of the anaerobic samples. Inside the glovebox,

    178 pellets were finely spread on glass plates, dried (25 °C, 24 h, in the dark) and grinded using a

    179 mortar and pestle. Vivianite can be found when samples are dried at room temperature, even in

    180 the presence of oxygen. Higher temperatures for sample drying should be avoided. Above 70 °C,

    181 in the presence oxygen, vivianite is transformed within hours into an amorphous iron phosphate

    182 compound (Čermáková et al., 2015). Thus, in such sludge samples, vivianite disappears (Poffet,

    183 2007).

    184 Samples for XRD analyses were filled in glass capillaries and sealed first with modelling clay and

    185 then superglue. Right before analyses, glass capillaries were sealed using a burner. The

    186 measurements were done on a PANalytical X´Pert PRO diffractometer with Cu-Ka radiation (5-

    187 80 o2θ, step size 0.008

    o). The results from XRD analyses were made semi-quantitative by

    188 determining the amorphous and crystalline peak area of the spectra (Origin Pro 9). This allows the

    189 determination of the degree of crystallinity and thus of the total mineral share of the sample. All

    190 samples that were analysed by Mössbauer spectroscopy were also analysed by XRD. In addition,

    191 two samples that were sampled in the aerated treatment lines (referred to as A-stage and Line 2

    192 activated sludge samples) were analysed using XRD.

    193 For Mössbauer analyses, samples were filled in plastic rings, sealed with Kapton tape and super

    194 glue and then wrapped in parafilm. It was expected that considerable amounts of P are bound in

    195 vivianite, thus a vivianite standard was prepared according to Roldan et al., 2002. Transmission

    196 57

    Fe Mössbauer spectra were collected at 4.2 and 300 K with conventional constant-acceleration

  • ACCEPTED MANUSCRIPT

    10

    197 and sinusoidal velocity spectrometers using a 57

    Co(Rh) source. Velocity calibration was carried

    198 out using an α-Fe foil. The Mössbauer spectra were fitted using Mosswinn 4.0 (Klencsár, 1997).

    199 Morphology and elemental compositions of sludge particles in the grinded solids was also

    200 analysed by SEM-EDX. Samples for SEM-EDX were exposed to air during measurements.

    201 Extractability of Fe in digested sludge was investigated using water, to extract water soluble Fe

    202 (pH=7, Wolf et al., 2009). Na-pyrophosphate solution (0.1 mol L-1

    , pH=9.5) was used to extract

    203 and quantify organic bound Fe. Pyrophosphate was used to extract organic bound Fe and Fe

    204 minerals mainly in soil but also from sewage sludge, vivianite was partially dissolved with this

    205 extract (Carliell-Marquet et al., 2009; McKeague, 1967; van Hullebusch et al., 2005). With

    206 pyrophosphate extraction no distinction between Fe(II)/Fe(III) could be made. Ammonium

    207 oxalate (0.2 mol L-1

    NH4-oxalate, pH=3) extracts poorly crystalline Fe, it was used to determine

    208 Fe(II)/Fe(III) in activated sludge before (Rasmussen and Nielsen, 1996). Each extraction was

    209 done in separate butyl rubber stoppered serum bottles, the extracts were added to wet sludge

    210 (n=3). Oxygen in the extracts was removed using headspace gas exchange equipment with a gas

    211 mixture containing 70 % N2 and 30 % CO2 in 5 cycles. The extract:TS ratios were 100:1 for H2O

    212 and pyrophosphate and 1000:1 for oxalate. All samples were shaken in the dark (16 h, 30 oC,

    213 100 rpm) before analysing Fe in the filtered (0.45 µm) but not centrifuged extracts.

    214 2.3 Estimate P bound to Fe

    215 P bound in vivianite was calculated from results of semi-quantitative XRD and Mössbauer

    216 spectroscopy. Additionally, to determine sulphide extractable P, 0.5 molar Na2S solution was

    217 added to 2 L digested sludge (molar Fe:P=0.55) and surplus sludge from Line 1 (molar

    218 Fe:P=0.25) from Leeuwarden and to digested sludge from Nieuwveer (molar Fe:P =0.73) in molar

    219 ratios S2-

    :Fe of 1.5. These samples were taken several months before/after the other samples. For

  • ACCEPTED MANUSCRIPT

    11

    220 Leeuwarden, molar Fe:P were very similar to the sludge used for other analyses, digested sludge

    221 showed an Fe:P=0.56 and Line 1 an Fe:P=0.28 but for Nieuwveer the sludge for the sulphide

    222 experiments had a lower Fe:P (0.73 vs 0.89). It was assumed that sulphide extracts specifically P

    223 bound to FeP (Kato et al., 2006). The experiments were done in a gastight reactor with pH control

    224 (pH=7.5) with a reaction time of at least 24 h. Samples from the reactor were taken using N2

    225 flushed syringes, filled in N2 rinsed plastic centrifuge tubes under a stream of N2 and centrifuged

    226 (15 min, 3200 G). Subsequently, sulphide, ortho-P (o-P) and the elemental composition were

    227 determined in samples that were filtered using N2 flushed syringes and filters (0.45 µm). At the

    228 end of these experiments at least 1 mmol sulphide L-1

    was still in solution indicating that the

    229 extraction was not sulphide limited. The maximum amounts of P that could be bound to Fe, Mg

    230 and Al were quantified by using the elemental composition of the TS (Table A. 1 & Table A. 2).

    231 For these calculations it was assumed that all solid Mg is present as struvite (molar Mg:P =1), all

    232 Fe as vivianite (molar Fe:P=1.5) and Al as a precipitate with a molar Al:P of 1.5 (Hsu, 1976).

    233 3 Results

    234 3.1 Mass balances

    235 In the STP Leeuwarden, mass balances showed, that the influent Fe load equals approximately the

    236 dosed Fe (Figure 1A). The effluent load was approximately 15 % of the influent Fe and 10 % of

    237 the influent P. The solid molar Fe:P ratio almost doubles from 0.33 before to 0.57 after anaerobic

    238 digestion due to digestion of external sludge with a high Fe (8.4 g Fe /kg sludge) and low P (not

    239 detectable) content. About 95 % of the external sludge originates from two cheese factories which

    240 use Fe(III) as flocculent.

  • ACCEPTED MANUSCRIPT

    12

    C

    US

    241

    242

    243

    244

    245

    246

    247 Figure 1: Daily mass balances for Fe and P in the STPs Leeuwarden (A) and Nieuwveer (B). Underlined 248 numbers were calculated (BP = belt press).

    249 For Nieuwveer, the mass balance showed that, dosed Fe is about three times the Fe entering via

    250 the influent (Figure 1B). The effluent load was 15 % of the influent Fe and 25 % of the influent P.

    251 The Fe and P loads from the A and B stage to the anaerobic digestion are similar. During

    252 digestion, the molar Fe:P ratio increased, due to external sludge, from 0.76 to 0.89.

    253 The mass balances were established by a single sampling campaign in Nieuwveer and three

    254 sampling campaigns in Leeuwarden. It was not intended to make a comprehensive mass balance

    255 which would require several samplings throughout the year. The mass balance served to identify

    256 main Fe sources and sinks in the STPs. For Nieuwveer, the calculated P loads of the influent,

    257 effluent and into the digester were about 20 % higher, the external sludge P input about 20 %

    258 lower when compared to the average yearly P loads for 2014 which were determined from daily P

    259 measurements on pooled samples (number from yearly balance/from our balance): influent (460

    260 vs 593 kg P d-1

    ), effluent (118 vs 153 kg P d-1

    ), external sludge (139 vs 110 kg P d-1

    ) and to the

    261 digester (411 vs 490 kg P d-1

    ). Since loads (except for the external sludge) were consistently

  • ACCEPTED MANUSCRIPT

    13

    262 higher for our measurements it can be assumed that patterns of Fe and P loads represent typical

    263 situations for the STP. For Leeuwarden, average yearly P loads in 2014 in the influent were also

    264 about 20 % higher (318 vs 264 kg P d-1

    ) and almost the same for the effluent (28 vs 26 kg P d-1

    ).

    265 Phosphorus flows into the digester are not regularly determined in Leeuwarden.

    266 The maximum gap for the mass balance was about 15 % for Fe in the STP Nieuwveer, mainly

    267 caused by an excess of Fe leaving the digester. In contrast, the gap in the P mass balance was only

    268 5 %. The gap in the Fe balance is most likely due to the lack of a representative external sludge

    269 sample. In Nieuwveer, external sludge is delivered in irregular intervals from various STPs

    270 applying CPR (Al/Fe dosing) and EBPR respectively. The external sludge sample in Nieuwveer

    271 was taken from a storage tank that, most likely, contained sludge also from a non Fe dosing plant.

    272 This explained why we underestimate Fe input into the digester whereas the P loads can be traced

    273 back.

    274 3.2 Dissolved Fe(II)&Fe(III)

    275 Fe(II)/Fe(III) stock solutions were added (n=3) to filtrates, obtained from digested sludge and

    276 from surplus sludge of Line 1 in Leeuwarden, to test the reliability of the ferrozine method (Table

    277 1). In filtrates from surplus sludge, Fe(II) was overestimated by about 7 % and Fe(III) by about

    278 4 %. In filtrates from digested sludge Fe(II) was added. Here, Fe(II) was underestimated by 2 %

    279 and Fe(III) overestimated by about 5 %. When Fe(III) was added it was overestimated by about 1

    280 %. These results indicate that the method can reliably detect dissolved Fe(II) and Fe(III) in

    281 sewage samples.

    282 After digestion, dissolved Fe in Leeuwarden sludge was surprisingly dominated by Fe(III), 1.6 mg

    283 L-1

    (Table 1). Also in digested sludge in Nieuwveer about half of the dissolved Fe was detected as

  • ACCEPTED MANUSCRIPT

    14

    284 Fe(III), 3.0 mg L-1

    . In general, dissolved Fe in most samples was dominated by Fe(III). This

    285 Fe(III) could be free Fe(III) or Fe(III) which was complexed by organic ligands such as humic

    286 substances (Table 1, Buffle, 1990).

    287 Although standard addition was successful, the results should, especially after digestion, be

    288 regarded with some caution. Fe(II)/Fe(III) were determined reliably, even in the presence of

    289 dissolved organic matter (DOM, 16–25 mg DOC L-1

    ) using the ferrozine method (Verschoor and

    290 Molot, 2013; Viollier et al., 2000). However, Viollier et al., 2000, added Fe(III) only. Verschoor

    291 and Molot, 2013 found Fe(II) and Fe(III) successfully back. Yet, when added, Fe(II) could be

    292 present as free Fe(II), whereas the Fe(II) that was already in the sample could partly also be

    293 present in complexed forms (Buffle, 1990). In surplus sludge, dissolved organic matter was on the

    294 same order of magnitude (Nieuwveer A-stage: 15 mg DOC L-1

    , Leeuwarden Line 1: 20 mg DOC

    295 L-1

    ) compared to the successful standard additions described before. In digested sludge, DOC

    296 concentrations were much higher, in Nieuwveer, 320 mg DOC L-1

    and in Leeuwarden, 126 mg

    297 DOC L-1

    .

    298 With the ferrozine assay, as we applied it, only free Fe(II) was detected (Jackson et al., 2012).

    299 When part of the Fe(II) was complexed by DOM, it was not detected in our first step, in which

    300 Fe(II) is quantified. Subsequently, to determine Fe(III), the sample pH was lowered, a reducing

    301 agent was added and the sample incubated (12 h). Under these conditions, complexed Fe(II) is

    302 mobilized and could be incorrectly assigned to Fe(III) (Gaffney et al., 2008; Jackson et al., 2012,

    303 Rasmussen and Nielsen, 1996). That also explains why total Fe levels from ICP-OES and from

    304 the ferrozine measurements matched very well. With ICP-OES free and complexed Fe is detected.

    305 To measure free and total Fe(II) in the samples the method of Gaffney et al., 2008 could be

    306 established for sewage samples. Additionally, a complementary method to determine Fe

  • ACCEPTED MANUSCRIPT

    15

    307 speciation e.g. by voltammetry would help to eliminate analytical uncertainties (Buffle, 1990). In

    308 Leeuwarden, samples from Lines 1 and 2 and from the influent were on-site filtered and directly

    309 added in Ferrozine to test if free Fe(II) is present, no colour reaction was visible.

    310 Despite all efforts, it cannot be excluded that part of the Fe(II) was oxidized during sampling or

    311 sample processing due to high sensitivity of Fe(II) to oxygen (Verschoor and Molot, 2013).

    312 Subsequently, total dissolved Fe concentrations may decrease due to precipitation of ferric iron

    313 oxides. Ferrous iron can even get microbial oxidized in absence of oxygen (Nielsen and Nielsen,

    314 1998). An opposing mechanism, that could occur after sampling and during sample transport, is

    315 the conversion of solid Fe(III) oxides to soluble Fe(II) by iron reducing bacteria (IRB).

    316 Accordingly, dissolved Fe(II) concentrations doubled within 24 h in samples from the A-stage and

    317 B-stage in Nieuwveer when they were incubated at 30 °C (data not shown).

    318 Classifying complexed Fe(II) as Fe(III) by the ferrozine assay and oxygen contamination, could

    319 explain the presence of dissolved Fe(III) after the anaerobic digestion. From a chemical point of

    320 view all dissolved Fe should be present as Fe(II). During anaerobic digestion, highly reducing

    321 conditions, including the formation of strong reducing agents like sulphide, prevail for more than

    322 three weeks. Also others found significant amounts of dissolved Fe(III) after similar periods of

    323 anaerobic incubation (Cheng et al., 2015). An increase of the oxidation-reduction potential over

    324 time could indicate that anaerobic conditions did not prevail in these experiments. In our

    325 discussion we will focus on total dissolved iron levels instead of the oxidation state of the

    326 dissolved Fe.

  • ACCEPTED MANUSCRIPT

    16

    327 Table 1: Dissolved Fe(II) and Fe(III) measurements from STPs Leeuwarden and Nieuwveer

    328

    329 3.3 Solids

    330 3.3.1 XRD

    331 XRD analyses revealed that vivianite and quartz were present in all samples (all XRD

    332 diffractograms and peak assignments are included in the supporting information). In the STP

    333 Leeuwarden, struvite was the dominating crystalline P phase (Table 2). During anaerobic

    334 digestion the relative share of struvite decreases compared to quartz and vivianite. In the STP

    335 Nieuwveer vivianite, as the only P containing crystalline phase, was detected in all samples. In

    336 digested solids and both A-stage samples from Nieuwveer, a peak at around 29.4 °2Th with

    337 intensities between 3.9 and 8.6 % could not be assigned.

    ID

    Fe(II)

    mg L-1

    Fe(III)

    mg L-1

    Fe (total)

    mg L-1

    Le

    eu

    wa

    rde

    n

    Surplus sludge Line 1, Fe(III) 0.1 0.5 0.6

    Activated sludge Line 2, Fe(II) 0.1 0.5 0.6

    Surplus sludge Line 2, Fe(II) 0.6 0.6 1.1

    Digested sludge 0.6 1.6 2.1

    Nie

    uw

    ve

    er A-stage: after FeII dosing 0.0 0.8 0.8

    A-stage: Surplus Sludge 18.3 12.7 31.0

    B-stage: Surplus Sludge 0.0 1.9 1.9

    Digested sludge 2.9 3.0 5.9

    Sta

    nd

    ard

    ad

    dit

    ion

    Filtrate (undigested) + Fe(II): 11.8 mg L-1

    12.7 (±0.24) 0 12.7

    Filtrate (undigested) + Fe(III): 11.1 mg L-1

    0 11.6 (±0.0) 11.6

    Filtrate (digested) + Fe(II): 11.3 mg L-1

    11.1 (±0.08) 0.6 (±0.14) 11.7

    Filtrate (digested) + Fe(III): 10 mg L-1

    0 10.1 (±0.5) 10.1

  • ACCEPTED MANUSCRIPT

    17

    T P E

    C C

    A

    338 Table 2: Results of semi quantitative XRD and VS analyses expressed as % of the total solids.

    339

    340 3.3.2 SEM-EDX

    341 In both STPs, no large particles with an overlap of Fe and P were found before the anaerobic

    342 digestion using SEM-EDX. Iron and P were homogenously distributed in the samples. After

    343 anaerobic digestion larger FeP particles (between 20 and 150 µm in diameter) with different

    344 crystalline morphologies were found (Figure 2). These particles showed Fe:P ratios between 1.1

    345 and 1.7 when measured by EDX.

    346

    347 Figure 2: SEM images of particles in digested sludge solids sampled in Leeuwarden (left) and Nieuwveer 348 (right). EDX showed Fe:P ratios between 1.1 (Leeuwarden) and 1.7 (Nieuwveer).

    Sampling station

    Quartz

    (%)

    Vivianite

    (%)

    Struvite

    (%)

    XRD

    amorphous

    (%)

    VS

    (%)

    Le

    eu

    wa

    rde

    n

    Line 1, Fe(III): Surplus sludge 7 2 11 80 70

    Line 2, Fe(II): Activated sludge 7 3 10 79 66

    Line 2, Fe(II): Surplus sludge 6 3 7 84 68

    Digested sludge 21 6 11 63 62

    Nie

    uw

    ve

    er A-stage: Activated Sludge 10 7 0 83 78

    A-stage: Surplus Sludge 8 6 0 86 80

    B-stage: Surplus Sludge 11 8 0 81 78

    Digested sludge 11 8 0 63 60

  • ACCEPTED MANUSCRIPT

    18

    349 3.3.3 Mössbauer spectroscopy

    350 Results of Mössbauer measurements at 4.2 K are summarized in Table 3 (all spectra and

    351 measurements at 300 K are included in the supporting information). Mössbauer spectroscopy at

    352 liquid helium temperature is more powerful as it reveals unambiguously the oxidation states and

    353 magnetic properties of the different Fe structures. The samples from Leeuwarden with digested

    354 solids and the vivianite standard showed signs of oxidation (25 to 28 % of vivianite was oxidized

    355 in the standard and in digested sludge respectively). Before measurements, these samples were

    356 sealed followed by storage at ambient atmosphere inside glass bottles with screw caps for about 1

    357 month. Subsequently, other samples were stored inside the glovebox until measurement and no

    358 signs of oxidation were visible, as indicated by the absence of oxidized vivianite.

    359 The spectra acquired with the vivianite standard (Figure 3) showed that about 75 % of the

    360 vivianite was not affected by oxidation and allowed to obtain a spectrum with parameters that are

    361 in good agreement with the literature for the two Fe(II) sites (Gonser and Grant, 1976). The

    362 oxidized magnetically split Fe(III) species in this standard might be an intermediate valence state

    363 between Fe(III) and Fe(II) like in magnetite (Harker and Pollard, 1993). Others suggested that

    364 oxidation of vivianite results in the formation of amorphous FeP (Miot et al., 2009), Lepidocrocite

    365 (Roldan et al., 2002) or lipscombite, beraunite or rockbridgite (Leavens, 1972).

  • ACCEPTED MANUSCRIPT

    19

    I R

    C

    US

    366

    367 Figure 3: Mössbauer spectra obtained at different temperatures with the vivianite standard.

    368 Samples taken from Line 1, Fe(III) dosing and Line 2, Fe(II) dosing in Leeuwarden were virtually

    369 the same. Between 94 and 96 % of the total Fe was Fe(II). Iron in vivianite represented 36 and 32

    370 % of the total Fe in Lines 1 and 2 respectively. The significant (33 and 35 %) paramagnetic

    371 contribution to the spectra which was not magnetically split at 4.2 K was assigned to Fe(II) in

    372 pyrite, FeS2. All other Fe species in these samples (summing up to about 30 %) could not be

    373 decisively assigned. Mössbauer spectra of a digested sludge sample taken in Leeuwarden to which

    374 sulphide was added contained the unknown Fe(II) compound that was still paramagnetic at 4.2 K

    375 (data not shown as this sample was exposed to oxygen), which contributed 20-21 % to the total Fe

    376 pool in the samples from Lines 1 and 2. Thus, we assumed this very well defined compound (Γ =

    377 0.4 mm s-1

    ) is a sulphur phase. However, many iron sulphide (FeSx) and iron sulphate compounds

    378 can be excluded as they are magnetic split at 4.2 K or because they have different Mössbauer

    379 parameters (Mullet et al., 2002; Sklute et al., 2015; Yoshida and Langouche, 2013). The presence

    380 of FeP minerals cannot be excluded (Dyar et al., 2014), for instance, the Mössbauer parameters of

  • ACCEPTED MANUSCRIPT

    20

    381 anapaite, Ca2Fe2+

    (PO4)2· 4H2O are close to the values we obtained (Eeckhout et al., 1999).

    382 Overall, we cannot assign this spectra to a certain Fe phase. The Fe(III) phase is an iron oxide

    383 possibly hematite, Fe2O3 (Murad and Cashion, 2004).

    384 The spectra obtained for the digested sludge sample in Leeuwarden showed that vivianite was the

    385 only FeP present (73 %). About 28 % of Fe comes from oxidized vivianite and 45 % of the Fe

    386 from vivianite unaffected by oxidation. The remaining 27 % of Fe(II) in this sample was pyrite.

    387 In the samples from the STP Nieuwveer, Fe(II) dominated as well. The samples contained

    388 vivianite, pyrite, an Fe(III) having Mössbauer parameters resembling those of hematite, Fe2O3

    389 (Murad and Cashion, 2004) and a paramagnetic (doublet) Fe(II) species that might be assigned to

    390 vivianite. The isomer shift of this Fe(II) is close to the one of vivianite, and the quadrupole

    391 splitting is also consistent with paramagnetic vivianite. Our measurements were made close to the

    392 Neel temperature (magnetic ordering temperature) of vivianite (12 K). It could be that some

    393 dispersed vivianite structures or vivianite structures with impurities are still paramagnetic at 4.2

    394 K. The quantification of vivianite using XRD suggested that this unknown Fe(II) is vivianite

    395 (Table 4). However, this Fe(II) could also be vivianite overlapping with another phase; as well as

    396 it can be another Fe compound other than vivianite. In the surplus sludge of the A-stage about 69

    397 % of the Fe was present as vivianite, additional 18 % as the potential vivianite phase, 9 % as

    398 pyrite and 4 % as Fe2O3. The Fe in the surplus sludge sampled from the B-stage was assigned to

    399 vivianite (55 %), to the potential vivianite phase (33 %), pyrite (7 %) and to Fe2O3 (5 %). In

    400 digested solids, 54 % of the Fe could be firmly assigned to vivianite, 27 % were assigned to the

    401 potential vivianite phase, the share of pyrite was (15 %) and the remaining Fe (4 %) was assigned

    402 to Fe2O3. For the subsequent discussions, it was assumed that the Fe(II) species that could not

    403 clearly be assigned to vivianite with Mössbauer spectroscopy was in fact vivianite.

  • ACCEPTED MANUSCRIPT

    21

    404 In all samples from Nieuwveer and in digested sludge from Leeuwarden no iron phosphate

    405 minerals besides vivianite were present. In surplus sludge from Leeuwarden the presence of other

    406 FeP phases than vivianite, cannot be excluded. Only, minor fractions of P can be adsorbed to the

    407 Fe2O3.

    408 Table 3: Results of Mössbauer measurements at 4.2 K. Experimental uncertainties: Isomer shift (IS): ± 409 0.01 mm s-1; Quadrupole splitting (QS): ± 0.01 mm s-1; Line width (Γ): ± 0.01 mm s-1; Hyperfine field: ± 410 0.1 T; Spectral contribution: ± 3%.

    Sample IS

    (mm·s-1

    )

    QS

    (mm·s-1

    ) Hyperfine field (T)

    Γ

    (mm·s-1

    ) Phase Spectral

    contribut

    ion (%)

    Leeuwarden

    Surplus Sludge Line 1: (Fe(III)

    dosing

    0.27 0.37

    0.93

    1.29

    1.01 1.26

    0.88 0.05

    - 2.47

    1.00 3.71

    - 51.8 19.3

    -

    11.4 24.8

    0.84 0.58

    0.74

    0.47

    1.07 1.07

    Fe2+

    Fe3+

    Fe2+

    Fe2+

    Fe2+

    Fe2+

    Pyrite

    Fe2O3

    Vivianite I

    Vivianite II

    33 6

    5

    20

    14 22

    Leeuwarden

    Surplus Sludge

    Line 2: (Fe(II) dosing)

    0.29 0.37

    1.07

    1.31 1.01 1.11

    0.84 -0.20

    -

    2.48 0.61 3.37

    - 51.4

    24.7

    - 11.4 26.4

    0.87 0.58

    0.77

    0.46 1.07 1.07

    Fe2+

    Fe3+

    Fe2+

    Fe2+

    Fe2+

    Fe2+

    Pyrite

    Fe2O3

    Vivianite I

    Vivianite II

    35 4

    8

    21 15 17

    Leeuwarden

    digested solids

    0.42

    0.50

    0.71 1.20

    1.25

    0.79

    -0.84

    0.81 0.50

    2.70

    -

    46.0

    46.3 10.0

    26.7

    0.72

    1.41

    1.41 1.18

    1.18

    Fe2+

    Fe3+

    Fe2+

    Fe2+

    Fe2+

    Pyrite

    Oxidized

    vivianite

    Vivianite I

    Vivianite II

    27

    15

    13 24

    21

    Nieuwveer

    Surplus sludge

    A-stage

    0.27

    0.38

    1.21

    1.22 1.21

    1.00

    -0.19

    2.79

    2.27 3.13

    -

    51.9

    - 15.0 26.4

    0.92

    0.86

    0.79

    0.64 0.64

    Fe2+

    Fe3+

    Fe2+

    Fe2+

    Fe2+

    Pyrite

    Fe2O3

    Dispersed vivianite

    Vivianite I

    Vivianite II

    9

    4

    18

    21 48

    Nieuwveer

    Surplus sludge

    B-stage

    0.33 0.37

    1.16

    1.26 1.18

    0.88 -0.14

    2.92

    2.23 3.32

    - 49.8

    -

    14.9 26.2

    0.87 0.86

    0.92

    0.93 0.93

    Fe2+

    Fe3+

    Fe2+

    Fe2+

    Fe2+

    Pyrite

    Fe2O3

    Dispersed vivianite

    Vivianite I

    Vivianite II

    7 5

    33

    18 37

    Nieuwveer

    Digested solids

    0.33 0.37

    1.21

    1.16 1.20

    0.88 0.20

    2.86

    2.21 3.05

    - 48.9

    -

    15.0 26.8

    0.87 0.86

    1.23

    0.77 0.77

    Fe2+

    Fe3+

    Fe2+

    Fe2+

    Fe2+

    Pyrite

    Fe2O3

    Dispersed vivianite

    Vivianite I

    Vivianite II

    15 4

    27

    15 39

    Vivianite

    Standard

    0.35

    0.77

    1.35

    1.36

    0.45

    -0.54

    2.36

    3.14

    44.2

    50.2

    14.8

    27.4

    1.24

    1.15

    0.64

    0.53

    Fe3+

    Fe2+

    Fe2+

    Fe2+

    Oxidized

    vivianite

    Vivianite I

    Vivianite II

    13

    12

    29

    46

    411

  • ACCEPTED MANUSCRIPT

    22

    412 3.3.4 Extractions

    413 Water, pyrophosphate and ammonium oxalate were used to extract Fe from digested sludge

    414 sampled in Leeuwarden (Figure 4A) and Nieuwveer (Figure 4B). Water was the mildest extract

    415 and dissolved 0.3 and 1 % of the total solid Fe in Leeuwarden and Nieuwveer respectively. This

    416 was expected, considering the relatively low solubility of FeSx and vivianite that dominated the

    417 digested sludge samples (Al-Borno and Tomson, 1994; Davison, 1991). However, in both STPs

    418 about 60 % of the water extractable Fe was Fe(III). During pyrophosphate extraction all dissolved

    419 Fe species were quantified, summing up to between 40 % (Nieuwveer) and 60 % (Leeuwarden) of

    420 the Fe. Considering the Mössbauer measurements, Fe bound in pyrite or vivianite must be part of

    421 this fraction. Thus, the Fe extracted using pyrophosphate was mainly of non-organic origin.

    422 Pyrophosphate extracts rather Fe from vivianite than from FeSx (Carliell-Marquet et al., 2009). In

    423 Leeuwarden, the Fe fraction in pyrophosphate (58±7 %) is in a similar range as vivianite (57 %,

    424 Mössbauer spectroscopy). However, in Nieuwveer, the pyrophosphate extracted Fe (41±1%) was

    425 much less than Fe bound in vivianite (81 %, Mössbauer spectroscopy). Ammonium oxalate

    426 extracted from Leeuwarden digested sludge all Fe (85 % as Fe(II)) and around 80 % of the Fe in

    427 Nieuwveer digested sludge (65 % as Fe(II)). Compared to the Mössbauer measurements, this

    428 indicates that the oxalate extraction and subsequent spectrophotometric determination of

    429 Fe(II)/Fe(III) overestimate the Fe(III) content by about 10 and 25 %. The described uncertainties

    430 for the spectrophotometric measurements for dissolved Fe(II)/Fe(III) could also affect the results

    431 of the Fe extraction.

  • ACCEPTED MANUSCRIPT

    23

    T IP R

    C S

    U

    AN

    M

    432 Figure 4: Extraction of Fe from digested sludge using different extracts. The error bars indicate standard 433 deviation (n=3).

    434 3.3.5 P bound to FeP

    435 The maximum amount of P bound that could be bound to Fe was estimated by using the elemental

    436 composition of the solids. Sulphide extraction was used to dissolve P bound to Fe. Phosphorus

    437 bound in vivianite was determined by semi-quantitative XRD and by Mössbauer spectroscopy

    438 (Table 4). The results for Leeuwarden indicate that in surplus sludge between 9 % (Mössbauer

    439 spectroscopy) and 13 % (XRD) and after digestion between 18 % (XRD) and 29 % (Mössbauer

    440 spectroscopy) of the P is bound in vivianite. According to XRD, the majority of P was bound in

    441 struvite in surplus (43 %) and digested sludge (35 %). These values are higher than the maximum

    442 values obtained from the elemental compositions . Thus, semi quantitative XRD overestimated the

    443 struvite content in the sludge.

  • ACCEPTED MANUSCRIPT

    24

    444 In the A-stage and the B-stage, estimates were in good agreement, about 50 % of the P in the A-

    445 stage and about 40 % of the P in the B-stage were bound in vivianite. After digestion, the

    446 estimates differ considerably between sulphide extraction (31 % of P bound to Fe) and Mössbauer

    447 spectroscopy and XRD (the latter two suggested 47 % and 53 % of the P are bound in vivianite).

    448 Additional, in Nieuwveer and Leeuwarden a maximum of 25 and 8 % of the total P could be

    449 bound to Al respectively.

    450 Table 4 indicates that, the elemental compositions of the samples tends to overestimate P bound to

    451 Fe as it does not take into account non Fe-FeP species (e.g. iron oxides, FeSx or organic bound

    452 Fe). The same principle applies in estimating P bound to Mg in struvite or to Al in aluminium

    453 phosphorus compounds (AlP). XRD may underestimate P bound to Fe as only the P bound to

    454 vivianite is detected. Also Mössbauer spectroscopy may underestimate P bound to Fe as not all

    455 compounds were identified and as the Fe2O3 can bind P.

    456 Table 4: Estimating P bound to FeP in different sewage (sludge) samples (n.d. = not determined).

    457

    458 4 Discussion

    459 Significant Fe loads entered both STPs via the influent, which could originate from the municipal

    460 sewage itself, from groundwater infiltration and from Fe dosing into the sewer system (Hvitved-

    Leeuwarden

    Average Line 1 & 2 Digested Sludge % of total P % of total P

    Vivianite/FeP Struvite/MgP Vivianite/FeP Struvite/MgP

    qXRD 13 43 18 35

    Mössbauer 9 - 29 -

    Elemental composition 26 36 36 25

    Sulfide 11 - 26 -

    Nieuwveer

    A-stage Surplus sludge B-stage Surplus Sludge Digested Sludge % of total P % of total P % of total P

    Vivianite/FeP Struvite/MgP Vivianite/FeP Struvite/MgP Vivianite/FeP Struvite/MgP

    qXRD 54 0 37 0 53 0

    Mössbauer 52 - 38 - 47 -

    Elemental composition 55 15 43 14 59 14

    Sulfide n.d. - n.d. - 31 -

  • ACCEPTED MANUSCRIPT

    25

    461 Jacobsen et al., 2013; van den Kerk, 2005). This often neglected, but nevertheless, large Fe input

    462 could assist in P removal in STPs (Gutierrez et al., 2010). Despite of its significant contribution,

    463 the speciation of the influent Fe and whether it can support CPR or not was not determined. The

    464 Fe dosing in both STPs (as for most other STPs in The Netherlands) was relatively low. The

    465 molar ratios of Fe dosed to P entering via the influent was in Leeuwarden 0.13 and in Nieuwveer

    466 0.42 (Figure 1). External Fe sources (i.e. influent and external sludge) contributed to about 80 %

    467 of the total Fe in the STP Leeuwarden. Here, a large input of Fe via the external sludge into the

    468 digester was identified. This suggests that Fe dosing can be significantly reduced, the external Fe

    469 input is sufficient to prevent H2S emissions during anaerobic digestion. In Nieuwveer, the dosed

    470 Fe contributed more significantly to the total Fe budget, yet still 35 % of the total Fe load

    471 originated from the inflowing sewage and from external sludge.

    472 Dissolved Fe was measured to identify equilibrium concentrations with Fe compounds. However,

    473 it turned out that during the dynamic conditions in the treatment lines of any STP (oxidizing and

    474 reducing conditions coupled to high microbial activities) measuring of a static/equilibrium Fe

    475 concentration is arbitrary. Thus, the reported dissolved Fe levels should be seen as an order of

    476 magnitude for these zones. Except for the influent and effluent samples, by far most of the Fe is

    477 part of the solid fraction. Accordingly, it was shown that even Fe(II), as product of IRB, can

    478 remain part of the solid phase (Rasmussen and Nielsen, 1996). The high dissolved Fe

    479 concentrations in the surplus sludge of the A-stage in Nieuwveer (about 30 mg Fe L-1

    ) highlighted

    480 the possibility of a slow or insufficient Fe(II) oxidation resulting in small dispersed Fe(III) and

    481 dissolved Fe(II). Sufficient oxidation and formation of Fe(III) oxides would cause a more rapid

    482 precipitation compared to Fe(II) (Ghassemi and Recht, 1971; Oikonomidis et al., 2010).

  • ACCEPTED MANUSCRIPT

    26

    483 Improving the aeration of Fe(II) or dosing of Fe(III) may help to improve the limited COD

    484 removal in the A-stage of this STP (De Graaff et al., 2015).

    485 The methodology which we employed for ammonium oxalate extraction gave only rough

    486 estimates about the Fe(II)/Fe(III) content in the sludge. The pyrophosphate extraction did not

    487 reliably extract organic Fe, as explained earlier (Stucki, 2013). In this study, Mössbauer

    488 spectroscopy was the most reliable method for quantifying and identifying Fe and FeP

    489 compounds. In contrast to XRD, Mössbauer spectroscopy can detect also amorphous Fe and FeP

    490 phases with very small particle sizes in low abundances provided appropriate standards have been

    491 prepared. On the other hand XRD detects all crystalline P compounds, also the ones that do not

    492 contain Fe. Mössbauer spectroscopy and XRD collectively showed that the solid Fe compounds of

    493 surplus sludge and anaerobic digested sludge were dominated by the ferrous phosphate mineral

    494 vivianite. Ferric iron did not play a significant role in any of the solid samples. Besides vivianite,

    495 the other major Fe compound was pyrite (Table 3).

    496 In a membrane bioreactor with anoxic/aerobic zones, Fe(III) dominated the solid iron pool (Wu et

    497 al., 2015) also in sludge sampled from the aeration tank of an STP using Fe(II) for CPR,

    498 ammonium oxalate extraction showed that Fe(III) dominated (Rasmussen and Nielsen, 1996).

    499 However, in our samples, regardless of aerobic zone in the STPs, Fe(II) was dominant.

    500 How is that possible? First, despite aerated areas, the sludge itself is partly non-aerated e.g. during

    501 low loading rates on weekends or in the night, in settlers and in the anoxic zones allowing the

    502 reduction of Fe(III). In flocs, oxygen free conditions can prevail throughout the treatment process

    503 due to diffusion limitation and when relatively low dissolved oxygen set-points are used. Thus,

    504 once vivianite is formed, anoxic conditions in flocs may help to channel it, without oxidation,

  • ACCEPTED MANUSCRIPT

    27

    505 through the aerated nitrification zone. Both, ours (SEM-EDX) and earlier research (Frossard et al.,

    506 1997; Zelibor et al., 1988) showed that vivianite is often part of an organic matrix.

    507 High activity of IRB in STPs has been measured which could result in rapid Fe(III) reduction and

    508 thus vivianite production (Cheng et al., 2015; Rasmussen and Nielsen, 1996). Assuming the

    509 reduction rates from Rasmussen and Nielsen, 1996, it would take between 19 h and 4 days in

    510 Leeuwarden and between 24 h and 5 days in Nieuwveer to reduce all solid Fe in the surplus

    511 sludge. These figures also indicate that Fe(III) reduction after sampling could influence the

    512 oxidation state of the Fe in samples. Once vivianite is formed, its chemical oxidation is relatively

    513 slow, on the time scale of weeks (Miot et al., 2009; Roldan et al., 2002). The oxidation by

    514 anaerobic nitrate-reducing iron-oxidizing bacteria was faster: it took approximately 16 days for

    515 complete oxidation (Miot et al., 2009). We could not find information on how long iron-oxidizing

    516 bacteria in the presence of oxygen would need for the oxidation of Fe(II) in vivianite.

    517 In Nieuwveer and in Line 2 in Leeuwarden, where Fe(II) is dosed for CPR, the mechanisms of

    518 vivianite formation were not obvious. Vivianite could either directly precipitate from solution or

    519 formed as a result of Fe(III) reduction. Indirect chemical Fe(III) reduction, induced by e.g.

    520 sulphide, FeSx or via humic substances (Biber et al., 1994; Golterman, 2001; Kappler et al., 2004)

    521 or direct Fe(III) reduction by IRB (Azam and Finneran, 2014; Cheng et al., 2015; Nielsen, 1996;

    522 Zhang, 2012) may have caused the formation of Fe(II) and subsequent precipitation of vivianite.

    523 Vivianite could also be precipitated directly from solution as a result of Fe(II) dosing, possibly

    524 combined with insufficient oxidation of Fe(II) (Ghassemi and Recht, 1971). In Line 1 in

    525 Leeuwarden where Fe(III) is dosed, also most of the solid Fe was present as Fe(II), mainly as

    526 vivianite. Here, chemical or biologically Fe(III) reduction must play a role. To what extent

    527 vivianite forms already in the sewer systems cannot be determined by our measurements.

  • ACCEPTED MANUSCRIPT

    28

    528 When Fe(III) is used for CPR, it was suggested that first Fe(III) oxides form which cause the P

    529 removal via co-precipitation or adsorption (Smith et al., 2008). If the Fe:P ratio of these initial

    530 ferric FeP is higher than the one of vivianite (molar Fe:P=1.5) as suggested before by Fulazzaky et

    531 al., 2014 and Luedecke et al., 1989, then Fe(III) reduction and subsequent formation of vivianite

    532 can act as a net sink for P. Hence, oxidation of Fe(II) in vivianite could result in P release due to a

    533 higher molar Fe:P ratio in the formed products (Miot et al., 2009; Roldan et al., 2002). In case of

    534 FeP with a molar Fe:P of 1 (e.g. strengite), Fe(III) reduction would cause a slight net P release.

    535 However, more significant P release could only be expected when vivianite formation would be

    536 prevented as documented in the presence of sulphide when FeSx are formed (Roden and Edmonds,

    537 1997). Accordingly, in our experiments, addition of sulphide to the vivianite containing sludge

    538 caused a relatively quick (4 h) and significant P release.

    539 Vivianite is very efficient in removing P from solution due to a very low solubility (pksp≈36, Al-

    540 Borno and Tomson, 1994). Fe(II) dosing for o-P removal in oxygen free conditions resulted in

    541 vivianite formation (Ghassemi, Recht 1971). The same researchers demonstrated that, in pure

    542 water, stoichiometry of o-P removal was more efficient with Fe(II) compared to Fe(III), resulting

    543 in lower residual o-P concentrations at optimum pH, and Fe(II) showed an optimum pH for o-P

    544 removal (pH=8) closer to common sewage. On the other hand, faster kinetics of Fe(III)P

    545 formation, faster settling of the formed Fe(III)P, a broader pH range for o-P removal and better

    546 COD flocculation properties were found for Fe(III) (Ghassemi and Recht, 1971; Gregory and

    547 O'Melia, 1989). In oxygen free freshwater (O'Connell et al., 2015; Rothe et al., 2014) and even in

    548 marine sediments (Jilbert and Slomp, 2013), in anoxic soils (Nanzyo et al., 2013; Peretyazhko and

    549 Sposito, 2005) and in septic tanks (Azam and Finneran, 2014) vivianite received attention as it

    550 plays an important role in P retention (see recent review by Rothe et al., 2016). For the formation

  • ACCEPTED MANUSCRIPT

    29

    551 of spherical vivianite in sediments a model has been suggested based on the presence of polymeric

    552 gel structures (Zelibor et al., 1988). At least one of the crystals we found in Nieuwveer (Figure

    553 2B) resembles the crystals produced by Zelibor et al., 1988, indicating that the mechanism of

    554 vivianite formation could be similar in sediments and in biological STPs.

    555 Similar to Frossard et al., 1997, with SEM-EDX we found larger crystals (up to 150 µm in

    556 diameter) with a Fe:P ratio close to the one of vivianite in digested sludge. Such large crystals

    557 were not found before digestion. The growth of the vivianite particles during digestion may be the

    558 result of the additional SRT of several weeks under constantly anaerobic conditions. Ostwaldt

    559 ripening, particle aggregation or crystal growth at elevated temperature in the digester may have

    560 caused the growth of vivianite particles/crystals. However, vivianite showed relatively slow

    561 crystal growth rates in pure solutions with higher vivianite supersaturations than observed in our

    562 samples (Madsen and Hansen, 2014) and vivianite is not stable in the presence of sulphide

    563 (Nriagu, 1972). Moreover, various inhibitors of Ostwald ripening, like DOC, are present during

    564 the digestion process. The apparent growth of vivianite particles during the digestion process is

    565 not yet fully understood.

    566 XRD could not detect crystalline FeSx in any of the samples analysed. Whereas Mössbauer

    567 spectroscopy revealed that pyrite contributed significantly to the solid Fe pool and even was

    568 present in the surplus sludge (9–33 % of the total Fe). The pyrite in these solids could originate

    569 from the sewer system or they were formed during the treatment process (Ingvorsen et al., 2003;

    570 Nielsen et al., 2005; van den Brand et al., 2015). Oxidation of FeSx in STPs is on a time scale of

    571 hours (Gutierrez et al., 2010; Ingvorsen et al., 2003; Nielsen et al., 2005), it could occur in aerated

    572 zones of the STPs. However, if located in the core of the sludge flocs, FeSx might, similar to

    573 vivianite, pass aerated zone without being oxidized.

  • ACCEPTED MANUSCRIPT

    30

    574 Quantification of P bound in FeP was performed by various approaches (Table 4). Vivianite

    575 bound P contributed in Leeuwarden about 10 % before digestion and around 30 % after digestion

    576 to the total solid P, according to Mössbauer spectroscopy. The quantifications using XRD

    577 suggested, P in struvite contributes before digestion around 43 % and after digestion about 35 %

    578 of the total P. This decrease can be explained by the external input of Fe in the digester. The

    579 dissolved P concentrations are usually quite high in anaerobic digesters due to organic matter

    580 degradation and, in EBPR plants, due to the release of polyphosphates from phosphate

    581 accumulating organisms. Thus, vivianite formation is not limited by the supply of P. The Fe from

    582 the external sludge will partly react with P to form vivianite, Fe dosing to anaerobic digesters is

    583 also a measure to prevent struvite scaling as vivianite is preferably formed (Mamais et al., 1994).

    584 Some of the added Fe could react with sulphide to form FeSx. Further P could be bound in

    585 biomass (phosphate accumulating organisms, cell material and debris) or in amorphous

    586 compounds associated with metals like Al, Mg or Ca. To be able to identify and quantify these P

    587 species would require the application of techniques like 31

    P-NMR, sequential extraction or X-ray

    588 absorption spectroscopy (e.g. Frossard et al., 1994; Wu et al., 2015). If, however, sufficient Fe(II)

    589 is available, vivianite is expected to be the dominant inorganic solid P compound in digesters.

    590 This would make a recovery technology targeting vivianite vastly more attractive. For Nieuwveer,

    591 Mössbauer spectroscopy indicates a decrease in vivianite bound P during digestion. Here the Fe:P

    592 increases only slightly due to external sludge input. Thus, the formation of FeSx during anaerobic

    593 digestion on expenses of vivianite causes a decrease in P bound to vivianite.

    594 XRD might not be able to detect small particles of vivianite and amorphous FeP which Mössbauer

    595 spectroscopy does detect. We consider XRD as a semi-quantitative method. In contrast to our

    596 expectations, XRD did not underestimate the vivianite bound P and results of Mössbauer

  • ACCEPTED MANUSCRIPT

    31

    597 spectroscopy were very similar (Table 4). This apparent match supports the assumption that the

    598 Fe(II) fraction in Nieuwveer, that Mössbauer spectroscopy could not clearly assign to vivianite, is

    599 actually vivianite. However, also the XRD results in Nieuwveer bear some uncertainty due to the

    600 presence of a peak that could not be assigned. Maximum quantities of P bound in FeP, AlP and

    601 MgP were estimated using the elemental composition of the TS. It was assumed that all Fe, Al and

    602 Mg is bound to P and thus other fractions of these elements were neglected. However, the

    603 elemental composition was, at least for the Fe, able to give good estimates on P bound to Fe.

    604 Sulphide was added to sample to extract P bound to Fe (Kato et al., 2006). In Leeuwarden, the

    605 sulphide extractable P fractions in digested sludge and in surplus sludge from Line 1 matched very

    606 well with P in vivianite obtained from Mössbauer spectroscopy. In Nieuwveer, the release of P

    607 from digested sludge in response to sulphide addition was much lower (31 %) compared to the P

    608 bound in vivianite (about 50 %). However, also the molar Fe:P was about 20 % lower in the

    609 sludge that was used for sulphide extraction. Translating the P release efficiency of sulphide to the

    610 sludge with the 20 % higher Fe:P ratio, we would expect a P release of about 40 %. Hence,

    611 sulphide extraction and Mössbauer spectroscopy would match better. It seems likely that the gap

    612 between sulphide extraction and the other methods is due to a difference in the sludge samples.

    613 Perhaps the released P, re-precipitated with other metals. However, from potential counter ions

    614 (Mg, Ca and Al), only Ca concentrations dropped noteworthy by 2 mmol L-1

    (net P release was 13

    615 mmol L-1

    ). Or else vivianite particles were present in Nieuwveer in another form (see Mössbauer

    616 results) then the ones in Leeuwarden (e.g. more crystalline, enclosed by other minerals/organic)

    617 that made vivianite less reactive/unreactive to sulphide exposure. Overall, all methods gave good

    618 estimates for P bound to Fe and for P bound in vivianite. The elemental composition is the easiest

    619 method but gives the less accurate result. Sulphide extractions is relatively simple, here P is

  • ACCEPTED MANUSCRIPT

    32

    620 released from all Fe compounds without determining the type of FeP present. XRD is a popular

    621 and common method. It allows the quantification of crystalline FeP only. In our case all Fe bound

    622 P was vivianite, hence the quantification worked well. In sludge with amorphous FeP or dispersed

    623 vivianite, XRD will not be able to quantify P bound to Fe. Mössbauer spectroscopy is, however,

    624 able to detect amorphous and crystalline Fe compounds very accurately. In Nieuwveer, about 1/3

    625 of the Fe(II) that was assigned to vivianite could also be another Fe(II) phase. Preparation of

    626 appropriate standards may help to identify this Fe(II) using Mössbauer spectroscopy in future. In

    627 general, Mössbauer gives very accurate qualitative/quantitative results but it should be used in

    628 combination with other complementary methods like XRD.

    629 5 Conclusion

    630 Mössbauer spectroscopy indicated that vivianite and pyrite were the dominating solid Fe

    631 compounds in the surplus and anaerobically digested sludge from two STPs applying CPR and

    632 EBPR. XRD confirmed that vivianite was the major FeP in the samples. None of the sludge

    633 samples contained a significant amount of Fe(III) although besides Fe(II) also Fe(III) was dosed.

    634 Likely, this is related to fast iron reduction processes and slow vivianite oxidation rates. Studying

    635 Fe chemistry, helped to identify measures on how sewage treatment can be improved. In

    636 Leeuwarden, Fe dosing, to prevent sulphide emissions, can be reduced. In Nieuwveer, improving

    637 the aeration to form Fe(III) would improve COD removal in the A-stage. To assess the role of

    638 vivianite and the potential of a P recovery technology targeting on FeP, further STPs with

    639 different treatment designs (higher Fe dosing and particularly higher Fe(III) dosing) should be

    640 analysed as well. If vivianite is a general iron precipitant in STPs it could offer new routes for P

    641 recovery.

  • ACCEPTED MANUSCRIPT

    33

    642 Acknowledgements

    643 This work was performed in the TTIW-cooperation framework of Wetsus, European Centre Of

    644 Excellence For Sustainable Water Technology (www.wetsus.nl). Wetsus is funded by the Dutch

    645 Ministry of Economic Affairs, the European Union Regional Development Fund, the Province of

    646 Fryslân, the City of Leeuwarden and the EZ/Kompas program of the “Samenwerkingsverband

    647 Noord-Nederland”. We thank the participants of the research theme “Phosphate Recovery” for

    648 their financial support and helpful discussions. We acknowledge the great support by employees

    649 of Waterschap Brabantse Delta, Wetterskip Fryslan and from the two STPs.

    650 6 References

    651 Al-Borno, A., Tomson, M.B., 1994. The temperature dependence of the solubility product

    652 constant of vivianite. Geochimica et Cosmochimica Acta 58 (24), 5373–5378.

    653 APHA, AWWA, WEF, 1998. Standard methods for the examination of water and wastewater,

    654 20th ed. 1998 ed. American Public Health Association, Washington, DC.

    655 Azam, H.M., Finneran, K.T., 2014. Fe(III) reduction-mediated phosphate removal as vivianite

    656 (Fe3(PO4)2⋅8H2O) in septic system wastewater. Chemosphere 97, 1–9.

    657 Biber, M.V., dos Santos Afonso, M., Stumm, W., 1994. The coordination chemistry of

    658 weathering: IV. Inhibition of the dissolution of oxide minerals. Geochimica et Cosmochimica

    659 Acta 58 (9), 1999–2010.

    660 Böhnke, B., 1977. Das Adsorptions-Belebungsverfahren. Korrespondenz Abwasser 24.

    661 Böhnke, B., Diering, B., Zuckut, S.W., 1997. Cost-effective wastewater treatment process for

    662 removal of organics and nutrients. Water Eng. Manag. 144 (7), 18–21.

  • ACCEPTED MANUSCRIPT

    34

    663 Buffle, J., 1990. Complexation reactions in aquatic systems: an analytical approach. Ellis

    664 Horwood series in analytical chemistry. Ellis Horwood.

    665 Carliell-Marquet, C., Oikonomidis, I., Wheatley, A., Smith, J., 2009. Inorganic profiles of

    666 chemical phosphorus removal sludge. Proceedings of the ICE - Water Management 163 (2),

    667 65–77.

    668 Carpenter, S.R., 2008. Phosphorus control is critical to mitigating eutrophication. Pro. Natl.

    669 Acad.Sci. USA 105, 11039–11040.

    670 Carpenter, S.R., Bennett, E.M., 2011. Reconsideration of the planetary boundary for phosphorus.

    671 Environ. Res. Lett. 6 (1), 14009.

    672 Čermáková, Z., Švarcová, S., Hradilová, J., Bezdička, P., Lančok, A., Vašutová, V., Blažek, J.,

    673 Hradil, D., 2015. Temperature-related degradation and colour changes of historic paintings

    674 containing vivianite. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy

    675 140, 101–110.

    676 Cheng, X., Chen, B., Cui, Y., Sun, D., Wang, X., 2015. Iron(III) reduction-induced phosphate

    677 precipitation during anaerobic digestion of waste activated sludge. Separation and Purification

    678 Technology 143, 6–11.

    679 Childers, D.L., Corman, J., Edwards, M., Else, J.J., 2011. Sustainability challenges of phosphorus

    680 and food: solutions from closing the human phosphorus cycle. Bioscience 61 (117-124).

    681 Cline, J.D., 1969. Spectrophotometric determination of hydrogensulfide in natural waters. Limnol.

    682 Oceangr. 14, 454–458.

    683 Davison, W., 1991. The solubility of iron sulphides in synthetic and natural waters at ambient

    684 temperature. Aquatic Science 53 (4), 309–329.

  • ACCEPTED MANUSCRIPT

    35

    685 De Graaff, M., Roest, D., Brand Van Den, T., Zandvoort, M., Duin, O., van Loosdrecht, M., 2015.

    686 Characterisation of high loaded wastewater treatment processes (A-stage) to increase energy

    687 production from wastewater: performance and design guidelines. Water research (accepted).

    688 De Ridder, M., De Jong, S., Polchar, J., Lingemann, S., 2012. Risks and opportunities in the

    689 global phosphate rock market: Robust strategies in times of uncertainty. Rapport / Centre for

    690 Strategic Studies no. 17 | 12 | 12. The Hague Centre for Strategic Studies, Den Haag.

    691 Dyar, M.D., Jawin, E.R., Breves, E., Marchand, G., Nelms, M., Lane, M.D., Mertzman, S.A.,

    692 Bish, D.L., Bishop, J.L., 2014. Mossbauer parameters of iron in phosphate minerals:

    693 Implications for interpretation of martian data. American Mineralogist 99 (5-6), 914–942.

    694 Eeckhout, S.G., Grave, E. de, Vochten, R., Blaton, N.M., 1999. Mössbauer effect study of

    695 anapaite, Ca 2 Fe 2+ (PO 4 ) 2 · 4H 2 O, and of its oxidation products. Physics and Chemistry

    696 of Minerals 26 (6), 506–512.

    697 El Samrani, A.G., Lartiges, B.S., Montarges-Pelletier, E., Kazpard, V., Barres, O., Ghanbaja, J.,

    698 2004. Clarification of municipal sewage with ferric chloride: The nature of coagulant species.

    699 Water research 38, 756–768.

    700 Frossard, E., Bauer, J.P., Lothe, F., 1997. Evidence of vivianite in FeSO4-flocculated sludges.

    701 Water research 31 (10), 2449–2454.

    702 Frossard, E., Tekely, P., Grimal, J.Y., 1994. Characterization of phosphate s ecies in urban sewage

    703 sludges by high-resolution solid-state 31P NMR. Eur J Soil Science 45, 403–408.

    704 Fulazzaky, M.A., Salim, N., Abdullah, N.H., Yusoff, A., Paul, E., 2014. Precipitation of iron-

    705 hydroxy-phosphate of added ferric iron from domestic wastewater by an alternating aerobic–

    706 anoxic process. Chemical Engineering Journal 253, 291–297.

  • ACCEPTED MANUSCRIPT

    36

    707 Gaffney, J.W., White, K.N., Boult, S., 2008. Oxidation State and Size of Fe Controlled by

    708 Organic Matter in Natural Waters. Environ. Sci. Technol. 42 (10), 3575–3581.

    709 Geraarts, B., Koetse, E., Loeffen, P., Reitsma, B., Gaillard, A., 2007. Fosfaatterugwinning uit

    710 ijzerarm slib van rioolwaterzuiveringsinrichtingen. STOWA, 83 pp.

    711 http://www.stowa.nl/Upload/publicaties2/mID\_4924\_cID\_3914\_74684671\_STOWA 2007

    712 31.pdf.

    713 Ghassemi, M., Recht, H.L., 1971. Phosphate Precipitation with Ferrous Iron. Water Pollution

    714 Control Research Series, 64 pp.

    715 Golterman, H.L., 2001. Phosphate release from anoxic sediments or ‘What did Mortimer really

    716 write?’. Hydrobiologia 450 (1/3), 99–106.

    717 Gonser, U., Grant, R.W., 1976. Determination of spin directions and electric fields gradient axes

    718 in vivianite by polarized recoil free y-Rays. Phys. Stat. Sol. 21 (331).

    719 Gregory, J., O'Melia, C.R., 1989. Fundamentals of flocculation. Critical Reviews in

    720 Environmental Control 19 (3), 185–230.

    721 Gutierrez, O., Park, D., Sharma, K.R., Yuan, Z., 2010. Iron salts dosage for sulfide control in

    722 sewers induces chemical phosphorus removal during wastewater treatment. Water research 44

    723 (11), 3467–3475.

    724 Harker, S.J., Pollard, R.J., 1993. A study of magnetite at 4.2 K and subject to strong applied

    725 magnetic fields. Nuclear Instruments and Methods in Physics Research Section B: Beam

    726 Interactions with Materials and Atoms 76 (1-4), 61–63.

    727 Higgins, M., Murthy, S., 2006. Understanding factors affecting polymer demand for thickening

    728 and dewatering. Water Environment Research Foundation; IWA Publishing, Alexandria, Va,

    729 London.

    http://www.stowa.nl/Upload/publicaties2/mID/_4924/_cID/_3914/_74684671/_STOWA

  • ACCEPTED MANUSCRIPT

    37

    730 Hsu, P.H., 1976. Comparison of iron(III) and aluminum in precipitation of phosphate from

    731 solution. Water research 10 (10), 903–907.

    732 Hvitved-Jacobsen, T., Vollertsen, J., Nielsen, A.H., 2013. Sewer processes: Microbial and

    733 chemical process engineering of sewer networks, 2nd ed. ed. CRC Press, Boca Raton.

    734 Ingvorsen, K., Nielsen, M.Y., Joulian, C., 2003. Kinetics of bacterial sulfate reduction in an

    735 activated sludge plant. Microb Ecol 46, 129^137.

    736 Jackson, A., Gaffney, J.W., Boult, S., 2012. Subsurface interactions of Fe(II) with humic acid or

    737 landfill leachate do not control subsequent iron(III) (hydr)oxide production at the surface.

    738 Environ. Sci. Technol. 46 (14), 7543–7550.

    739 Jetten, M., Horn, S., van Loosdrecht, M.C.M., 1997. Towards a more sustainable municipal

    740 wastewater treatment system. Water Science & Technology 35 (9), 171–180.

    741 Jilbert, T., Slomp, C.P., 2013. Iron and manganese shuttles control the formation of authigenic

    742 phosphorus minerals in the euxinic basins of the Baltic Sea. Geochimica et Cosmochimica

    743 Acta 107, 155–169.

    744 Kappler, A., Benz, M., Schink, B., Brune, A., 2004. Electron shuttling via humic acids in

    745 microbial iron(III) reduction in a freshwater sediment. FEMS Microbiology Ecology 47 (1),

    746 85–92.

    747 Kato, F., Kitakoji, H., Oshita, K., Takaoka, M., Takeda, N., Matsumoto, T., 2006. Extraction

    748 efficiency of phosphate from pre-coagulated sludge with NaHS. Water Science & Technology

    749 54 (5), 119.

    750 Klencsár, Z., 1997. Mössbauer spectrum analysis by Evolution Algorithm. Nuclear Instruments

    751 and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 129

    752 (4), 527–533.

  • ACCEPTED MANUSCRIPT

    38

    753 Kraal, P., Slomp, C.P., Forster, A., Kuypers, M., Sluijs, A., 2009. Pyrite oxidation during sample

    754 storage determinesphosphorus fractionation in carbonate-poor anoxic sediments. Geochimica

    755 et Cosmochimica Acta 73, 3277–3290.

    756 Kracht, O., Gresch, M., Gujer, W., 2007. A Stable Isotope Approach for the Quantification of

    757 Sewer Infiltration. Environ. Sci. Technol. 41 (16), 5839–5845.

    758 Leavens, P.B. (Ed.), 1972. Oxidation of vivianite in New Jersey Cretacerous greensands.

    759 Li, J., 2005. Effects of Fe(III) on floc characteristics of activated sludge. J. Chem. Technol.

    760 Biotechnol. 80 (3), 313–319.

    761 Luedecke, C., Hermanowicz, S.W., Jenkins, D., 1989. Precipitation of ferric phosphate in

    762 activated-sludge - A chemical model and its verification. Water Science & Technology 21 (4-

    763 5), 325–337.

    764 Macdonald, G.K., Bennett, E.M., Potter, P.A., Ramankutty, N., 2011. Agronomic phosphorus

    765 imbalances across the world’s croplands 108 (7), 3086–3091.

    766 Madsen, H., Hansen, H., 2014. Kinetics of crystal growth of vivianite, Fe3(PO4)2 x 8H2O, from

    767 solution at 25, 35 and 45 °C. Journal of Crystal Growth (401), 82–86.

    768 Mamais, D., Pitt, P.A., Cheng, Y.W., Loiacono, J., Jenkins, D., 1994. Determination of ferric

    769 chloride dose to control struvite precipitation in anaerobic sludge digesters. Water Environ Res

    770 66 (7), 912–918.

    771 McKeague, J.A., 1967. An evaluation of 0.1 M pyrophosphate and pyrophosphate-dithionite in

    772 comparison with oxalate as extractants of accumulation products in podzols and some other

    773 soils. Can. J. Soil. Sci. 47 (2), 95-&.

  • ACCEPTED MANUSCRIPT

    39

    774 Miot, J., Benzerara, K., Morin, G., Bernard, S., Beyssac, O., Larquet, E., Kappler, A., Guyot, F.,

    775 2009. Transformation of vivianite by anaerobic nitrate-reducing iron-oxidizing bacteria.

    776 Geobiology 7 (3), 373–384.

    777 Mullet, M., Boursiquot, S., Abdelmoula, M., Génin, J.M., Ehrhardt, J.J., 2002. Surface chemistry

    778 and structural properties of mackinawite prepared by reaction of sulfide ions with metallic

    779 iron. Geochimica et Cosmochimica Acta 66 (5), 829–836.

    780 Murad, E., Cashion, J., 2004. Mössbauer Spectroscopy of Environmental Materials and Their

    781 Industrial Utilization. Springer US, Boston, MA.

    782 Nanzyo, M., Onodera, H., Hasegawa, E., Ito, K., Kanno, H., 2013. Formation and Dissolution of

    783 Vivianite in Paddy Field Soil. Soil Science Society of America Journal 77 (4), 1452.

    784 Nielsen, A.H., Lens, P., Vollertsen, J., Hvitved-Jacobsen, T., 2005. Sulfide–iron interactions in

    785 domestic wastewater from a gravity sewer. Water research 39 (12), 2747–2755.

    786 Nielsen, J.L., Nielsen, P.H., 1998. Microbial Nitrate-DependentOxidation of Ferrous Iron in

    787 Activated Sludge. Environ. Sci. Technol. 32, 3556–3561.

    788 Nielsen, P., 1996. The significance of microbial Fe(III) reduction in the activated sludge process.

    789 Water Science & Technology 34 (5-6), 129–136.

    790 Nriagu, J.O., 1972. Stability of vivianite and ion-pair formation in the system fe3(PO4)2-

    791 H3PO4H3PO4-H2o. Geochimica et Cosmochimica Acta 36 (4), 459–470.

    792 O'Connell, D.W., Jensen, M.M., Jakobsen, R., Thamdrup, B., Andersen, T.J., Kovacs, A., Hansen,

    793 H., 2015. Vivianite formation and its role in phosphorus retention in Lake Ørn, Denmark.

    794 Chemical Geology 409, 42–53.

    795 Oikonomidis, I., Burrows, L.J., Carliell-Marquet, C., 2010. Mode of action of ferric and ferrous

    796 iron salts in activated sludge. J. Chem. Technol. Biotechnol. 85 (8), 1067–1076.

  • ACCEPTED MANUSCRIPT

    40

    797 Paul, E., Laval, M.L., Sperandio, M., 2001. Excess sludge production and costs due to phosphorus

    798 removal. Environmental technology 22, 1363–1371.

    799 Peretyazhko, T., Sposito, G., 2005. Iron(III) reduction and phosphorous solubilization in humid

    800 tropical forest soils. Geochimica et Cosmochimica Acta 69 (14), 3643–3652.

    801 Poffet, M.S., 2007. Thermal runaway of the dried sewage sludge in the storage tanks: from

    802 molecular origins to technical measures of smouldering fire prevention. Dissertation thesis.

    803 Rasmussen, H., Nielsen, P., 1996. Iron reduction in activated sludge measured with different

    804 extraction techniques. Water research 30 (3), 551–558.

    805 Recht, H.L., Ghassemi, M., 1970. Kinetics and mechanism of precipitation and nature of the

    806 precipitate obtained in phosphate removal from wastewater using aluminum(III) and iron(III)

    807 salts. Water Pollution Control Research Series. U.S. Dept. of the Interior, FWQA.

    808 Roden, E.E., Edmonds, J.W., 1997. Phosphate mobilization in iron-rich anaerobic sediments:

    809 Microbial Fe(III) oxide reduction versus iron-sulfide formation. Archiv für Hydrobiologie 139

    810 (3), 347–378.

    811 Roldan, R., Barron, V., Torrent, J., 2002. Experimental alteration of vivianite to lepidocrocite in a

    812 calcareous medium. clay miner 37 (4), 709–718.

    813 Rothe, M., Frederichs, T., Eder, M., Kleeberg, A., Hupfer, M., 2014. Evidence for vivianite

    814 formation and its contribution to long-term phosphorus retention in a recent lake sediment: A

    815 novel analytical approach. Biogeosciences 11 (18), 5169–5180.

    816 Rothe, M., Kleeberg, A., Hupfer, M., 2016. The occurrence, identification and environmental

    817 relevance of vivianite in waterlogged soils and aquatic sediments. Earth-Science Reviews 158,

    818 51–64.

  • ACCEPTED MANUSCRIPT

    41

    819 Scholz, R.W., Wellmer, F.W., 2016. Comment on: "Recent revisions of phosphate rock reserves

    820 and resources: a critique" by Edixhoven et al. (2014) – clarifying comments and thoughts on

    821 key conceptions, conclusions and interpretation to allow for sustainable action. Earth Syst.

    822 Dynam. 7 (1), 103–117.

    823 Seitz, M.A., Riedner, R.J., Malhotra, S.K., Kipp, R.J., 1973. Iron-phosphate compound

    824 identification in sewage sludge residue. Environ. Sci. Technol. 7 (4), 354–357.

    825 Siegrist, H., Salzgeber, D., Eugster, J., Joss, A., 2008. Anammox brings WWTP closer to energy

    826 autarky due to increased biogas production and reduced aeration energy for N-removal. Water

    827 Science & Technology 57 (3), 383–388.

    828 Singer, P.C., 1972. Anaerobic control of phosphate by ferrous iron: Anaerobic control of

    829 phosphate by ferrous iron. Journal Water Pollution Control Federation 44 (4), 663-&.

    830 Sklute, E.C., Jensen, H.B., Rogers, A.D., Reeder, R.J., 2015. Morphological, structural, and

    831 spectral characteristics of amorphous iron sulfates. J. Geophys. Res. Planets 120 (4), 809–830.

    832 Smith, S., Takacs, I., Murthy, S., Daigger, G.T., Szabo, A., 2008. Phosphate complexation model

    833 and its implications for chemical phosphorus removal. Water Environ Res 80 (5), 428–438.

    834 Stucki, J.W., 2013. Iron in soils and clay minerals. Springer, [Place of publication not identified].

    835 Szabó, A., Takács, I., Murthy, S., Daigger, G.T., Licskó, I., Smith, S., 2008. Significance of

    836 Design and Operational Variables in Chemical Phosphorus Removal. Water Environ Res 80

    837 (5), 407–416.

    838 Takács, I., Murthy, S., Smith, S., McGrath, M., 2006. Chemical phosphorus removal to extremely

    839 low levels: experience of two plants in the Washington, DC area. Water Science &

    840 Technology 53 (12), 21.

  • ACCEPTED MANUSCRIPT

    42

    841 Tchobanoglous, G., Burton, F.L., Stensel, H.D., 2013. Wastewater engineering: Treatment and

    842 reuse, 5th ed. ed. McGraw-Hill Higher Education; McGraw-Hill [distributor], New York,

    843 London.

    844 Thomas, A.E., 1965. Phosphat-Elimination in der Belebtschlammanlage von Männedorf und

    845 Phosphat-Fixation in See- und Klärschlamm. Vierteljahrschr. Naturforsch. Ges. Zürich 110,

    846 419–434.

    847 van den Brand, T. P. H., Roest, K., Chen, G. H., Brdjanovic, D., van Loosdrecht, M. C. M., 2015.

    848 Occu